• Nie Znaleziono Wyników

Uncertainty reduction and sampling efficiency in slope designs using 3D conditional random fields

N/A
N/A
Protected

Academic year: 2021

Share "Uncertainty reduction and sampling efficiency in slope designs using 3D conditional random fields"

Copied!
43
0
0

Pełen tekst

(1)

Delft University of Technology

Uncertainty reduction and sampling efficiency in slope designs using 3D conditional

random fields

Li, Y.; Hicks, M. A.; Vardon, P. J. DOI

10.1016/j.compgeo.2016.05.027 Publication date

2016

Document Version

Accepted author manuscript Published in

Computers and Geotechnics

Citation (APA)

Li, Y., Hicks, M. A., & Vardon, P. J. (2016). Uncertainty reduction and sampling efficiency in slope designs using 3D conditional random fields. Computers and Geotechnics, 79, 159-172.

https://doi.org/10.1016/j.compgeo.2016.05.027 Important note

To cite this publication, please use the final published version (if applicable). Please check the document version above.

Copyright

Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy

Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim.

This work is downloaded from Delft University of Technology.

(2)

Uncertainty reduction and sampling efficiency in slope designs using 3D

1

conditional random fields

2

Y.J. Li, M.A. Hicks, P.J. Vardon

3

Section of Geo-Engineering, Department of Geoscience & Engineering, Faculty of Civil Engineering

4

& Geosciences, Delft University of Technology, Delft, The Netherlands

5

Email: y.li-4@tudelft.nl; m.a.hicks@tudelft.nl; p.j.vardon@tudelft.nl

6

Cite as:

7

Li, Y. J., Hicks, M. A., & Vardon, P. J. (2016). Uncertainty reduction and sampling

8

efficiency in slope designs using 3D conditional random fields. Computers and

9

Geotechnics, 79: 159-172.

doi:10.1016/j.compgeo.2016.05.027

10

free link until 16 August, 2016. http://authors.elsevier.com/a/1THNi,63b~copS

11

This is the Authors’ final manuscript. (page number 1-27) 12

Abstract

13

A method of combining 3D Kriging for geotechnical sampling schemes with an existing random field 14

generator is presented and validated. Conditional random fields of soil heterogeneity are then linked 15

with finite elements, within a Monte Carlo framework, to investigate optimum sampling locations and 16

the cost-effective design of a slope. The results clearly demonstrate the potential of 3D conditional 17

simulation in directing exploration programmes and designing cost saving structures; that is, by 18

reducing uncertainty and improving the confidence in a project’s success. Moreover, for the problems 19

analysed, an optimal sampling distance of half the horizontal scale of fluctuation was identified. 20

Key words: conditional random fields, Kriging, reliability, sampling efficiency, spatial variability, 21

uncertainty reduction 22

(3)

1. Introduction

23

Soil properties exhibit three dimensional spatial variability (i.e. heterogeneity). In geotechnical 24

engineering, a site investigation may be carried out, and the data collected and processed in a 25

statistical way to characterise the variability [1–10]. The outcomes of the statistical treatment, e.g. the 26

mean property value, the standard deviation or coefficient of variation, and the spatial correlation 27

distance, may be used as input to a geotechnical model capable of dealing with the spatial variation 28

(e.g. a random field simulation). However, when it comes to making use of the field data, there arises 29

the question: How can we make best use of the available data? The idea is to use the data more 30

effectively, so that it is worth the effort or cost spent in carrying out the investigation, as well as the 31

additional effort in post-processing the data. The aim of this paper is to contribute towards answering 32

this question. 33

For example, cone penetration tests (CPTs) are often carried out in geotechnical field investigations, in 34

order to obtain data used in implementing the design of a structure. The amount of data from CPT 35

measurements is often larger than from conventional laboratory tests. This is useful, as a large 36

database is needed to accurately estimate the spatial correlation structure of a soil property. For 37

example, Fenton [3] used a database of CPT profiles from Oslo to estimate the correlation statistics in 38

the vertical direction, and Jaksa et al. [5] used a database from Adelaide to estimate the correlation 39

distances in both the vertical and horizontal directions. 40

In geotechnical engineering, a substantial amount of numerical work has been done using idealised 2D 41

simulations based on collected in-situ data (e.g. [4]), although a 3D simulation would be preferable 42

due to site data generally being collected from a 3D space. However, there are relatively few studies 43

simulating the effect of 3D heterogeneity due to the high computational requirements. Examples 44

include the effect of heterogeneity on shallow foundation settlement [11–13], on steady state seepage 45

[14–16], on seismic liquefaction [17] and on slope reliability [18–27]. 46

The above investigations all used random fields to represent the soil spatial variability and the finite 47

element method to analyse geotechnical performance within a Monte Carlo framework, a form of 48

(4)

analysis sometimes referred to as the random finite element method (RFEM) [28]. However, they did 49

not make use of the spatial distribution of related measurement data to constrain the random fields. In 50

other words, for those applications that are based on real field data, many realisations not complying 51

with the field data at the measurement locations will be included in the simulation, which, in turn, will 52

result in an exaggerated range of responses in the analysis of geotechnical performance. 53

Studies on conditional simulations are available in geostatistics in the field of reservoir engineering 54

[29]. However, there are not many studies dealing with soil spatial variability in geotechnical 55

engineering that utilise conditional simulation (some 2D exceptions include, e.g., [6, 30–32]). This is 56

partly due to the smaller amount of data generally available in geotechnical engineering, and partly 57

due to there often not being a computer program specially implemented for those situations where 58

there are sufficient data (e.g. CPT, vane shear test (VST)), especially in 3D. However, unconditional 59

random fields can easily be conditioned to the known measurements by Kriging [29, 33]. Hence, 60

following the previous 2D work of Van den Eijnden and Hicks [31] and Lloret-Cabot et al. [30], this 61

paper seeks to implement and apply conditional simulation in three dimensional space, in order to 62

reduce uncertainty in the field where CPT measurements are carried out. 63

Usually, site investigation plans are designed to follow some regular pattern. For example, a 64

systematic grid of sample locations is generally used, due to its simplicity to implement [5]. Moreover, 65

although there are various sampling plans in terms of layouts, it is found that systematically ordered 66

spatial samples are superior in terms of the quality of estimates at unsampled locations [34]. Therefore, 67

this paper will be devoted to implementing a 3D Kriging algorithm for sampling schemes following a 68

regular grid. This will then be combined with an existing 3D random field generator to implement a 69

conditional simulator. However, extension to irregular sampling patterns is straightforward based on 70

the presented framework. 71

The implemented approach has been applied to two idealised slope stability examples. The first 72

demonstrates how the approach may be used to identify the best locations to conduct borehole testing, 73

and thereby allow an increased confidence in a project’s success or failure to be obtained. While it is 74

(5)

very important to pay sufficient attention to the required intensity of a site investigation (i.e. the 75

optimal number of boreholes) with respect to the site-specific spatial variability, as highlighted by 76

Jaksa et al. [12], the first example starts by focusing on the optimum locations for carrying out site 77

investigations for a given number of boreholes, before moving on to consider the intensity of testing. 78

The second example compares different candidate slope designs, in order to choose the best (most 79

cost-effective) design satisfying the reliability requirements. 80

For simplicity, this paper focuses on applications involving only a single soil layer (i.e. a single layer 81

characterised by a statistically homogeneous undrained shear strength), although the extension to 82

multiple soil layers is straightforward. Moreover, the effect of random variation in the boundary 83

locations between different soil layers can also be easily incorporated by conditioning to known 84

boundary locations (e.g. corresponding to where the CPTs have been carried out). 85

2. Theory and Implementation

86

2.1 Conditioning

87

A conditional random field, which preserves the known values at the measurement locations, can be 88

formed from three different fields [28, 35–36]: 89

(

)

( )

( )

( )

( )

rc ru km ks

Z

x

=

Z

x

+

Z

x

Z

x

(1) 90

wherexdenotes a location in space,Zrc( )x is the conditionally simulated random field,Zru( )x is the 91

unconditional random field,Zkm( )x is the Kriged field based on measured values at xi(i=1, 2,2,N), 92

( ) ks

Z x is the Kriged field based on unconditionally (or randomly) simulated values at the same 93

positions xi(i=1, 2,2,N), and N is the number of measurement locations. 94

The unconditional random field can be simulated via several methods [37]; for example, interpolated 95

autocorrelation [38], covariance matrix decomposition, discrete Fourier transform or Fast Fourier 96

transform, turning bands, local average subdivision (LAS), and Karhunen–Loeve expansion [39], 97

among others. The LAS method [40] is used in this paper. The Kriged fields are obtained by Kriging 98

(6)

[41], which has found extensive usage in geostatistics [42–43]. The LAS and Kriging methods are 99

briefly reviewed in the following sections. 100

2.2 Anisotropic random field generation using 3D LAS

101

The LAS method [40, 44] is used herein to generate the unconditional random fields, using statistics 102

(i.e. mean, variance and correlation structure) based on the observed field data. The LAS method 103

proceeds in a recursive fashion, by progressively subdividing the initial domain into smaller cells, until 104

the random process is represented by a series of local averages. The major advantage is its ability to 105

produce random fields of local averages whose statistics are consistent with the field resolution; that is, 106

it maintains a constant mean over all levels of subdivision, and ensures reduced variances as a function 107

of cell size based on variance reduction theory [45], taking account of spatial correlations between 108

local averages within each level and across levels. 109

The following covariance function is used in the subdivision process: 110

( )

(

)

2 1 2 2 3 2 1 2 3 1 2 3

2

2

2

,

,

exp

C

C

τ τ τ

σ

τ

τ

τ

θ

θ

θ

=

=

+ 

τ

(2) 111

where

σ

2is the variance of the soil property,

τ

is the lag vector, and θ1, θ2 and θ3, and τ1, τ2 and 112

τ3 are the respective scales of fluctuation and lag distances in the vertical and two lateral coordinate 113

directions, respectively. Herein, an isotropic random field is initially generated by setting θ1 = θ2 = θ3 114

= θiso; i.e. so that θiso equals the horizontal scale of fluctuation, θh. This field is then squashed in the 115

vertical direction to give the target vertical scale of fluctuation, θv. The 3D LAS implementation of 116

Spencer [25] has been used in this paper, and the reader is referred to Spencer [25] and Hicks and 117

Spencer [19] for more details. Note also that a truncated normal distribution has been used to describe 118

the pointwise variation in material properties [19]. 119

2.3 Kriging

120

In contrast to conventional deterministic interpolation techniques, such as moving least squares and 121

the radial point interpolation method, Kriging incorporates the variogram (or covariance) into the 122

(7)

interpolation procedure; specifically, information on the spatial correlation of the measured points is 123

used to calculate the weights. Moreover, standard errors of the estimation can also be obtained, 124

indicating the reliability of the estimation and the accuracy of the prediction. Kriging is a method of 125

interpolation for which the interpolated values are modelled by a Gaussian process governed by prior 126

covariances and for which confidence intervals can be derived. While interpolation methods based on 127

other criteria need not yield the most likely intermediate values, Kriging provides a best linear 128

unbiased prediction of the soil properties (Z) between known data [43, 46] by assuming the stationarity 129

of the mean and of the spatial covariances, or variograms. A brief review is first given to facilitate 130

understanding of the implementation. 131

Suppose that Z Z1, 2,2,ZNare observations of the random field

Z x

( )

at points x x1, 2,2,xN(i.e. 132

( ) ( 1, 2, , )

i i

Z =Z x i= 2 N ). The best linear unbiased estimation (i.e. Zˆ) of the soil property at some 133 locationx0is given by 134 0 0 1 1

ˆ ( )

(

) ( )

N N i i i i i i

Z

λ

Z

λ

Z

= =

=

=

x

x

x

(3) 135

in which N denotes the total number of observations and

λ

idenotes the unknown weighting factor 136

associated with observation point xi, which needs to be determined. 137

The weights in equation (3), for the estimation at any location

x

0, can be found by minimising the 138

variance (

σ

e2) of the Kriging error Zˆ−Z , which is given as 139

(

)

(

)

(

)

(

)

(

(

)

)

(

(

)

)

(

)

(

) (

)

2 2 2 2 2 0 00 1 1 1 2 2 2 0 0 0 1 1 1 0 0 0 1 1 1

ˆ

ˆ

var

(

) 2

(

) (

)

2

2

N N N e i j ij i i i j i N N N i j i j i i i j i N N N i j i j i i i j i

Z

Z

E

Z

Z

c

c

c

C

C

C

σ

λ λ

σ

λ

σ

σ

λ λ

σ

λ

σ

σ

λ λ γ

λ γ

γ

= = = = = = = = =

=

=

=

+

=

+

= −

+

∑∑

∑∑

∑∑

x

x

x

x

x

x

x

x

x

x

x

x

(4) 140 6

(8)

where var() denotes the variance operator and E[] is the expectation operator, cij =C

(

xixj

)

is the 141

covariance between ( )Z x and i Z x( j),ci0=C

(

xix0

)

is the covariance between ( )Z x and i Z x( 0), 142

and c00=C

(

x0x0

)

=C(0) =σ2 is the variance of 0

( )

Z x , which is estimated at the target location 143

0

x . The rearrangement in equation (4) makes use of the relationship between a variogram

γ

( )

τ and a 144

covariance function C

( )

τ ( i.e.

γ

( )

τ =C

( )

0 −C

( )

τ =

σ

2−C

( )

τ ) and the condition

1 1 N i i λ = =

(in 145

order to ensure that the estimator is unbiased, i.e. E Z

(

ˆ−Z

)

= , the weights must sum to one). 0 146

To minimise the error variance (i.e. equation (4)), the Lagrange method is used [43]. The weights can 147

then be found by solving the system of equations of size N + 1, for a constant mean: 148 1 1 1 1 1 0 1 0

(

)

(

)

1

(

)

(

)

(

)

1

(

)

1

1

0

1

N N N N N N

γ

γ

λ

γ

γ

γ

λ

γ

µ



 



 



 

=



 



 



 

x

x

x

x

x

x

x

x

x

x

x

x

(5) 149

in which

µ

is the Lagrangian parameter. For a mean following some trend, the modification to 150

equation (5) is straightforward and interested readers are referred to Fenton [46]. 151

Equation (5) may be expressed as 152

lhs x = rhs

γ λ γ (6) 153

Once equation (5) is solved, the estimated error variance can be expressed by 154

(

)

(

)

T 2 1 0 1 N e i i lhs rhs rhs i

σ

µ

λ γ

− =

= +

x

x

=

γ

γ

γ

(7) 155

where

λ

iis a function of the relative positioning of pointsxiand x0. 156

Note that the left-hand-side matrix γlhsis a function of only the observation point locations and 157

covariance between them. Therefore, it only needs to be inverted once, and then equations (5) and (3) 158

(9)

used repeatedly for building up the field of best estimates at different locations in space. In contrast, 159

the right-hand-side vector γrhschanges as a function of the spatial point x0, resulting in different 160

weight vectors λxthat are used in equation (3) to get the estimates (point by point) in the domain of 161

interest. 162

In geotechnical engineering, a sampling strategy following some pattern is generally adopted [1]. For 163

example, CPT sampling is often planned in the form of a regular grid on the ground surface [5]. It is 164

therefore desirable to implement the above Kriging algorithm in the context of some sampling design 165

with a regular pattern. While it is straightforward to implement in 2D, it is less so when implemented 166

in 3D. The most fundamental part is how the left-hand-side matrix of equation (6) is formed. The 167

authors have implemented 3D Kriging for the regular grid sampling strategy shown in Figure 1. The 168

way to set up the left-hand-side matrix and the right-hand-side vector is presented in the Appendix. 169

2.4 Computational efficiency

170

There are two aspects involved in the computational efficiency of the above Kriging implementation. 171

One is the total number of equations, which depends on the total number of data points (N = k×m×n, 172

where k and m are the number of CPT rows in the x and y directions respectively, and n is the number 173

of data points for each CPT profile, see Figure 1) contributing to the left-hand-side matrix; the other is 174

the number of points in the field (nf = nx×ny×nz, where nx, ny and nz are the number of points in the 175

three Cartesian directions) that need to be Kriged (i.e. how many times the algorithm will need to be 176

repeated, except for inverting the left-hand-side matrix). The higher the required field resolution (nf) 177

and the greater the total number of known data points (N), the longer the Kriging will take. In the case 178

of the CPT arrangement in Fig. 1, the size of matrix γlhs (see equation (5) or equation (A1)) and the 179

size of vector γrhs(see equation (5) or equation (A3)) in 3D are m2 and m times larger than those in 2D 180

(i.e. a cross-section in the x–z plane) respectively. The time it takes to Krige a full 3D field depends on 181

the processing time of each individual step and the number of times each step has to be performed. To 182

Krige a field of size nf, conditional to N measurement points, the total time may be approximated by 183

(10)

2 3

1 2

( , f) f

t N nc n N +c N (8) 184

where the first term represents the time needed for solving the system of equations for all field points 185

(i.e. nf times) (O(N 2

)) and the second represents the time needed for inverting the matrix

γ

lhs (i.e. only 186

once) (O(N3)). The constants c1 and c2 are functions of the CPU speed and the operation, and in this 187

case are in a ratio of approximately 4:1. Additionally, in all practical cases, nf >> N, so that the 188

calculation time depends mainly on the first term in the above equation; that is, on the number of times 189

(i.e. nf times) that the matrix–vector multiplication operation,

1 x lhs rhs

=

λ

γ

γ

, needs to be performed. 190

Note that nf = nx×ny×nz in 3D is ny times nf = nx×nz in 2D and N = k×m×n in 3D is m times N = k×n in 191

2D. In the examples reported in Section 4, all problems investigated are very long in the third 192

dimension compared to the cross-section. Therefore, the time consumed in a 3D analysis is 193

theoretically ny×m 2

times that of a 2D analysis, when neglecting the relatively fast, one-off matrix 194

inversion operation and other computation overheads, such as reading/writing and memory operations. 195

However, despite the significantly greater run-time requirements for Kriging in 3D (as compared to 196

2D), it is still far less than the time consumed in a nonlinear finite element analysis where plasticity 197

iterations are needed. For Example 1 in Section 4, where nx = 20, ny = 100 and nz = 20, it took, in serial 198

and on average, 134 hours in total for the 500 finite element analyses forming each Monte Carlo 199

simulation (3.0 GHz CPU), whereas Kriging 500 times took about 2.4 hours. In contrast, 500 Kriging 200

interpolations for a 2D cross-section analysis took approximately 8.5 seconds. It is noted that the 201

computation time used for Kriging 500 times is significant in comparison with a single finite element 202

analysis, and therefore should not be considered a pre-processing step if utilising parallel computation 203

for the finite element analyses. Therefore, the computing strategy developed to carry out the analyses 204

for Examples 1 and 2 in Section 4 (comprising around 30,000 realisations in total, and involving 205

30,000 3D Kriging interpolations) was to run the analyses in parallel (each Kriging and finite element 206

analysis serially on a single computation node) on the Dutch national grid e-infrastructure with high 207

performance computing clusters. 208

(11)

Note that it is possible to prescribe an appropriate neighbourhood size in the algorithm to reduce the 209

computational burden for 3D Kriging. For example, a neighbourhood size of 5×7×n may be used to 210

construct the left-hand-side matrix (see the neighbourhood denoted as a rectangle in Figure A.1(a), i.e. 211

by using only the nearest 4 CPT profiles). That is, only those CPT profiles that have a significant 212

influence (i.e. a lag distance within the range of the scale of fluctuation in equation (2)) on the point to 213

be estimated are used to construct the left-hand-side (LHS) matrix. However, using this strategy, for 214

each point (or each subset of points) to be estimated, the left-hand-side matrix is different and will 215

need to be inverted accordingly, so this could increase the computational time if there are a large 216

number of points or cells to be estimated. Therefore, a choice has to be made, to make sure that the 217

time saved by inverting a smaller matrix, instead of a bigger one, outweighs the time consumed by 218

inverting the left-hand-side matrices for all the (subgroups of) cells to be estimated for the case in 219

which a neighbourhood is used. And, of course, there is a trade-off between the estimation accuracy 220

and time saved when such a neighbourhood approach is used. The accuracy will increase as more 221

available data are used to do the Kriging estimation, and so the neighbourhood size depends on the 222

required accuracy and the scales of fluctuation. 223

Due to the relatively fast inversion of the LHS matrix in the current investigation (the maximum size 224

investigated is N = 500), all CPT profiles have been used for the Kriging in the examples in Section 4. 225

However, one neighbourhood strategy was investigated by using the 4 nearest CPT profiles, and the 226

following uncertainty reduction ratio (a 3D extension to the 1D definition in [31]) has been used to 227

assess the approximation error: 228

(

)

1 1 1 , , y x n z n n e i j k x y z i j k u n n n σ σ = = = =

∑∑∑

(9) 229

The approximation error may be evaluated by 230 n a u a u u E u − = (10) 231 10

(12)

where un and ua are the uncertainty reduction ratios when using a neighbourhood and when all CPT 232

profiles have been used, respectively. 233

One of the sampling strategies from Example 1 (Section 4, Fig. 9(b)) was used to evaluate the 234

approximation error and the results are listed in Table 1. It can be seen that using a neighbourhood of 235

the 4 nearest CPT profiles has been sufficient in this case. 236

3. Validation

237

The conditional simulation of a 5 m high (z), 5 m wide (x) and 25 m long (y) clay block, characterised 238

by a spatially varying undrained shear strength, is presented in this section to demonstrate the 239

procedure and the validity of the implementation described in Sections 2.1–2.3 (and the Appendix). 240

The idea is to show how the measured values are honoured, and to check whether or not the statistical 241

properties (e.g. covariance) of the random fields are maintained after conditioning. 242

The block is discretised into 20×100×20 cubic cells, with each cell of dimension 0.25 m. The mean of 243

the undrained shear strength is 40 kPa, and the standard deviation is 8 kPa. The degree of anisotropy 244

of the heterogeneity is ξ = 3, in which ξ = θhv and θv = 1.0 m. Five CPT measurement locations in 245

the y direction (at x = 2.5 m) are available, each comprising n = 20 data points at 0.25 m spacing in the 246

vertical direction. These ‘measured’ data have been obtained by sampling from a single independent 247

realisation of the spatial variability (i.e. representing the ‘actual’ in-situ variability). The interval 248

distance between the CPTs in the horizontal direction is Δy = 5 m, and the first CPT is located at y = 249

2.5 m. 250

Figure 2 shows an example realisation, to illustrate the stages involved in constructing the conditional 251

random field. It shows (a) the unconditional field generated using LAS, (b) the Kriged field based on 252

the unconditionally simulated cell values at the measurement locations, (c) the Kriged field based on 253

the measured data (taken from the reference field), and (d) the conditional random field. It can be seen 254

that the conditional field eliminates unrealistic values from the unconditional simulation by honouring 255

the measurement data at the measurement locations (e.g. corresponding to the centre of the dashed 256

circle in the case of the first CPT). The cross-section from which the CPTs were taken is also shown in 257

(13)

Figures 2(e) and 2(f), together with the known CPT profiles. It is seen that the known CPT profiles are 258

honoured in the conditional random field. Note that, in order to better visualise the fields, a local 259

colour scale is used for all sub-figures in Figure 2. 260

In order to validate the consistency of the conditioning, the following estimator of the correlation 261

structure along the vertical or horizontal directions of the random field is used to back-figure the 262 covariance structure: 263

(

)

(

)

1

1

ˆ

(

)

n j

ˆ

ˆ

j i Z i j Z i

C

j

Z

Z

n

j

τ

τ

µ

+

µ

=

= ∆ =

(11) 264

where j = 0, 1, …, n-1, n is the number of data points in the vertical or horizontal direction,

τ

jis the 265

lag distance between xi and

x

i j+ , ∆ is the distance between two adjacent cells vertically or τ 266

horizontally,

µ

ˆZis the estimated mean, Z is the random soil property and Zi is the sample of Z. The 267

correlation function is then ρ τˆ ( )j =Cˆ( )τj Cˆ(0), where Cˆ(0)=σˆ2Zand σˆ2Zis the estimated variance 268

[3]. 269

Figure 3 shows the back-figured (a) vertical and (b) horizontal covariances for the unconditional and 270

conditional random fields averaged over 200 realisations, as well as the sample (i.e. CPT) covariances 271

and exact covariances (i.e. equation (2) with only those terms that are associated with the vertical or 272

horizontal direction in 1D). It can be seen that the conditional random field preserves the covariance 273

structure reasonably well in both the horizontal and vertical directions, and that the correlation 274

function fits well the sample correlation for the first quarter of the data points (i.e. n/4) [5, 9]. It is also 275

seen that the covariance for the conditional field lies in between those for the unconditional field and 276

the sampling points. 277

4. Applications

278

Two simple examples concerning slope stability are presented in this section, to illustrate how the 279

technique presented in this paper may be used as an aid to geotechnical design. The first involves 280

finding the optimum locations for CPT profiles, in order to minimise the uncertainty in assessing the 281

(14)

reliability of a slope. The second involves a cost-effective design with regard to the slope angle when 282

field measurements have already been made (i.e. the positions where the CPT data were taken are 283

already known). 284

Both examples are presented in terms of the uncertainties in the slope response (with respect to factor 285

of safety). The factors of safety are calculated by 3D finite elements using the strength reduction 286

method [47], with the analyses being undertaken within a probabilistic (RFEM) framework; a 287

flowchart for carrying out such a simulation is shown in Figure 4. The undrained clay behaviour has 288

been modelled using a linear elastic, perfectly plastic Tresca soil model. The clay has a unit weight of 289

20 kN/m3, a Young’s modulus of 100 MPa and a Poisson’s ratio of 0.3. With reference to Figures 5 290

and 13, the finite element boundary conditions are: a fixed base, rollers on the back of the domain 291

preventing displacements perpendicular to the back face, and rollers on the two ends of the domain, 292

allowing only settlements and preventing movements in the other two directions (i.e. the out-of-slope-293

face and longitudinal directions). A full explanation of these boundary conditions is given in Spencer 294

[25] and Hicks and Spencer [19]. 295

The random field cell values are mapped onto the 2×2×2 Gauss points in each 20-node finite element, 296

in order to simulate the spatial variability more accurately [19, 48]. Note that the random fields (both 297

conditional and unconditional) have been mapped onto a finite element mesh with an element aspect 298

ratio equal to 2.0 (see Figure 5) to save time for the finite element analyses [25]. A detailed description 299

of how the random field cell values, in this case based on a cell size of 0.25×0.25×0.25 m, are mapped 300

onto the larger non-cubic finite elements is given in Hicks and Spencer [19]. 301

Note that field test (e.g. CPT) data are not directly used in the following examples. That is, the direct 302

measurements from geotechnical tests are typically not directly applicable in a design. Instead, a 303

transformation model is needed to relate the test measurement (e.g. tip resistance from a CPT test) to 304

an appropriate design property (e.g. the undrained shear strength) [49]. The uncertainty involved in the 305

transformation model is not considered in this paper. 306

(15)

4.1 Example 1

307

The first example considers a proposed 45˚, 5 m high, 50 m long slope, that is to be cut from a 308

heterogeneous clay deposit characterised by an undrained shear strength with the following statistics: 309

mean, µ = 20 kPa; standard deviation, σ = 4 kPa; vertical scale of fluctuation, θv = 1.0 m; and 310

horizontal scale of fluctuation, θh = 6.0 m. A question arises as to how to design the sampling strategy 311

for the soil deposit. For example, if 5 CPTs are to be conducted in a straight line along the axis of the 312

proposed slope, where is the best location to site the CPTs such that the designed slope will have the 313

smallest uncertainty in the realised factor of safety F? Hence, this example first investigates the 314

influence of the CPT locations on the standard deviation of the realised factor of safety, followed by 315

the influence of CPT intensity. 316

Figure 5 shows a cross-section through the slope, and 10 possible positions to locate the CPTs (i = 0, 317

1, …, 9). Note that the CPTs are taken to be equally spaced (i.e. at 10 m centres) in the third 318

dimension, and that the first and fifth CPTs are located at 5 m and 45 m along the slope axis (see 319

Figure 9(a)). Furthermore, the CPTs are carried out before the slope is excavated, in a block of soil of 320

dimensions 10×50×5 m as indicated in the figure. 321

Both conditional and unconditional RFEM simulations were carried out, using 500 realisations per 322

simulation, to investigate how the structure response (in this case, the realised factor of safety) 323

changes as the conditioning location changes. Figure 6(a) shows that the uncertainty in the realised 324

factor of safety reduces after conditioning, i.e. after making use of the available CPT information 325

about the soil variability, as indicated by the narrower distribution of realised factor of safety for the 326

conditional simulation. In this figure, the reduction in uncertainty is due to CPT data being taken from 327

location i = 5. 328

Figure 6(b) shows the sampling efficiency indices with respect to the different CPT locations, in which 329

the sampling efficiency index is defined as 330 u se i

I

σ

σ

=

(12) 331 14

(16)

where

σ

uis the standard deviation of the realised factor of safety for the unconditional simulation, and 332

i

σ

is the standard deviation of the realised factor of safety for the conditional simulation based on 333

column position i. Hence Ise=1 if the simulation is not conditioned. Clearly, there exists an optimum 334

position (in this case, i = 5) to locate the CPTs; i.e. the uncertainty is a minimum if the CPTs are 335

located along the crest of the proposed slope. In contrast, when i = 0 and i = 1, there is little 336

improvement, because the potential failure planes (in the various realisations) generally pass through 337

zones where the shear strength is, at most, only weakly correlated to values at the left-hand boundary 338

(due to θh being only 6 m in this case). It is interesting to note that, although there is not much 339

information included in the slope stability calculation when i = 9, i.e. for the CPTs at the slope toe, the 340

reduction in uncertainty is still noticeable, due to the CPTs being located in the zone where slope 341

failure is likely to initiate. This observation highlights that the location of additional information may 342

matter more than how much additional information there is (e.g. contrast the large difference in the 343

amount of directly utilised data between CPT locations i = 0 and i = 9). 344

However, it should be remembered that Figures 6(a)-6(b) are for the case of ξ = 6 (corresponding to θh 345

= 6 m) and that ξ often takes a larger value in practice. Figures 6(c)-6(f) show that, for ξ = 12 and ξ = 346

24, the reduction in uncertainty relative to the unconditional case is greater. Moreover, improved 347

values of Ise are obtained for CPT locations near the right and left boundaries, due to the higher 348

correlation of soil properties in the horizontal direction. 349

Figure 7 summarises the results as a function of the degree of anisotropy of the heterogeneity

ξ

. It is 350

seen that the best locations for carrying out the 5 CPTs are at i = 5, 6 and 7. As the value of

ξ

351

increases, the sampling efficiency indices increase due to the decreasing Kriging variance 2 e

σ

, as 352

illustrated in Figure 8 for a y–z slice at i = 5 (i.e. corresponding to where the CPTs are located). It is 353

seen that, for larger values of

ξ

, the Kriging variance between CPTs can drop well below the input 354

variance of the shear strength (i.e.

σ

e2

16 kPa

2). Moreover, carrying out CPTs at some distance to 355

the left or right of the slope crest for higher values of

ξ

can have a similar effect to carrying out CPTs 356

(17)

near the crest for smaller values of

ξ

. For example, Figure 7 shows that the sampling efficiency index 357

for

ξ

=24 at i= is approximately the same as that for 2

ξ

=12 at i = 5, 6 and 7. 358

Note that the same reference 3D random field is used to represent the ‘real’ field situation in 359

conditioning the random fields in each RFEM analysis. The 3D random fields are conditioned before 360

being mapped onto the finite element mesh, so that they are consistent with sampling the ground 361

before the slope is cut. Hence, for i = 6, 7, 8 and 9, although the CPT measurements are directly used 362

for fewer cells in the FE mesh, they nevertheless have an impact on all cell values via the lateral 363

spatial correlation of soil properties in the original ground profile. 364

If a second row of CPT tests (at position j) is to be performed in a second phase of the site 365

investigation (e.g. as illustrated in Figure 9(b)), the above procedure can be repeated by changing j in 366

the range 0–9 to locate the best positions for the new CPTs, assuming that the position of the first set 367

of CPT profiles has been set to i = 5. This is shown in Figure 10 for the case of

ξ

=6. Figure 10(a) 368

shows the probability distributions of the realised factor of safety for the unconditional simulation, the 369

conditional simulation for one row of CPTs at i = 5 and the conditional simulation for an additional 370

row of CPTs at position j = 0. It is seen that the confidence level in the project has been further 371

increased by the second phase of site investigation. Figure 10(b) shows the sampling efficiency indices 372

for various locations j of the second row of CPTs. It suggests that the best location for carrying out the 373

second phase of site investigation can be at either side of the slope crest (at a distance of 374

approximately 3 m (i.e. θh/2) from the crest). 375

To further investigate the influence of CPT intensity on the uncertainty in the realised factor of safety, 376

conditional simulations involving different numbers of CPTs (and thereby different distances (Δ) 377

between adjacent CPTs) have been carried out for the case of ξ = 6, 12 and 24. Figure 11 shows the 378

plan views of CPT layouts for ncpt = 3, 5, 9, 17 and 25 (corresponding to CPT spacings of Δ = 20, 10, 5, 379

3 and 2 m, respectively), with the locations of the CPTs in the x-direction being fixed at i = 5. Figure 380

12 shows the influence of CPT intensity on the sampling efficiency indices for the three values of

ξ

. 381

It is seen that there is only a marginal benefit in increasing the scope of the investigation by having 382

(18)

CPT spacings less than Δ ≈ θh/2, especially for the

ξ

=6 and

ξ

=12cases. For

ξ

=24, the sampling 383

efficiency index is as high as 4 when Δ ≈ θh/2, although more CPTs (i.e. Δ ≈ θh/4, ncpt = 9) may 384

improve the sampling efficiency to a value of 4.5. However, the general finding from Figures 10(b) 385

and 12 is that the optimal sampling distance is around θh/2 for the problem investigated, based on the 386

assumed correlation function. 387

4.2 Example 2

388

In the second example, a soil deposit characterised by spatially varying undrained shear strength is to 389

be excavated to form a slope of a certain angle. Site investigations have been conducted based on CPT 390

tests. The question is: In order to satisfy a target reliability level of, for example, 95%, as suggested in 391

Eurocode [50] and discussed in Hicks and Nuttall [51], how steep should the slope be designed? 392

Figure 13 shows three possible slope angles, with the corresponding finite element mesh 393

discretisations. The slope is 5 m high and 50 m long in the third dimension, and the left-hand boundary 394

is taken to be 15 m from the slope toe. Five CPTs were taken along the length of the slope at 10 m 395

centres, at the location of the column of Gauss points nearest the slope crest for the 1:1 slope, as seen 396

in the figure. The clay soil has a mean undrained shear strength of 21 kPa, a coefficient of variation of 397

0.2, a vertical scale of fluctuation of 1 m and a horizontal scale of fluctuation of 12 m. 398

The three candidate slopes are (vertical:horizontal) 1:2, 1:1 and 2:1. Based on only the mean undrained 399

shear strength, these three slopes have deterministic factors of safety Fd of 1.73, 1.29 and 1.07. Both 400

conditional and unconditional simulations were carried out to investigate the reliability of each slope, 401

and, for each simulation, 500 realisations were analysed. Note that, as in the previous example, one 402

reference random field is generated first and assumed to represent the real field situation. The 403

conditional random fields used in the RFEM analyses are therefore based on CPT measurements taken 404

from this ‘real’ field. 405

The stability of the slopes was calculated by the strength reduction method by applying gravitational 406

loading. The probability density functions of the realised factor of safety are shown in Figure 14 for 407

the three slopes, for both conditional and unconditional simulations. The deterministic factors of safety 408

(19)

Fd, i.e. the factors of safety based on the mean property values, are also shown. It is seen that, if 409

unconditional simulation is used, there is a significant chance that the 2:1 slope will fail (the 410

probability of failure is the area under the pdf for the realised factor of safety smaller than 1.0). 411

Unsurprisingly, the gentlest (i.e. 1:2) slope has the lowest probability of failure. However, once again, 412

conditional simulations significantly reduce the uncertainty in the structural response, as clearly 413

demonstrated by the narrower probability distributions. In particular, the reliability of the steepest 414

slope increases from 77% to 99% when the CPT measurements are taken into account. 415

The results show that, if unconditional simulations are used, the 1:1 and 1:2 slopes satisfy a target 416

reliability level of 95%, whereas the 2:1 slope does not. However, when the additional information 417

from the CPT profiles is used, all three cases meet the target reliability. This means that the 418

embankment may be designed to a slope angle of 2:1 if the CPT measurements are used in the 419

simulation, which is, if possible, a more logical thing to do. This has implications for the soil volume 420

to be excavated and thereby cost, although the cost can be site and situation dependent (e.g. on 421

whether there are nearby structures). A best design is a design that meets the requirements set by 422

standards, while, at the same time, minimising the cost. In this case, the steepest slope is likely to be 423

the most cost-effective design. 424

5. Conclusions

425

An approach for conditioning 3D random fields based on CPT measurements has been implemented 426

and validated, and then applied to two numerical examples to illustrate its potential use for 427

geotechnical site exploration and cost-effective design. It has been shown that conditional simulations 428

based on CPT data are able to increase the confidence in a design’s success or failure. Indeed, the 429

reliability from a conditional simulation can be thought of as a conditional reliability (or conditional 430

probability of failure not occurring), i.e. based on a ‘posterior’ distribution of the structure 431

performance after taking account of the spatial distribution of all the measured CPT data points. In 432

contrast, the unconditional simulation based on random field theory only results in a ‘prior’ 433

distribution of the structure response. This was clearly demonstrated by the updating of the probability 434

(20)

density distributions in the two numerical examples. Although Bayesian updating is not used in this 435

paper, the effect is similar. 436

If further CPT measurements are required, the approach can be repeated for updating the response 437

probability density function. In this way, the confidence in the probability of failure or survival will be 438

further increased. In fact, in many cases a multi-stage site investigation may be carried out, with the 439

results of the initial analysis guiding further field tests. As demonstrated in the first example, if a 440

second stage of site exploration were to be conducted, it is possible to find out the optimum location 441

for the additional testing. This highlights the method’s potential use in directing site exploration 442

programmes and thereby improving the efficient use of field measurements. For the first example 443

considered in this paper, an optimal sampling distance of half the horizontal scale of fluctuation was 444

identified when an exponential correlation function is used. For the second example, the conditional 445

simulation led to a more cost-effective design. 446

Acknowledgements

447

This research was funded by the China Scholarship Council (CSC) and by the Section of Geo-448

Engineering at Delft University of Technology. It was carried out on the Dutch National e-449

infrastructure with the support of the SURF Foundation. Special thanks are given to SURFsara advisor 450

Anatoli Danezi for her kind support in developing a computing strategy. 451

References

452

[1] DeGroot DJ, Baecher GB. Estimating autocovariance of in-situ soil properties. ASCE Journal of

453

Geotechnical Engineering 1993; 119 (1): 147–66.

454

[2] Fenton GA. Estimation for stochastic soil models. ASCE J. Geotech. Geoenv. Eng. 1999; 125(6): 470–85.

455

[3] Fenton GA. Random field modeling of CPT data. ASCE J. Geotech. Geoenv. Eng. 1999; 125(6): 486–98.

456

[4] Hicks MA, Onisiphorou C. Stochastic evaluation of static liquefaction in a predominantly dilative sand fill.

457

Géotechnique 2005; 55(2): 123–33.

458

[5] Jaksa MB, Kaggwa WS, Brooker PI. Experimental evaluation of the scale of fluctuation of a stiff clay. In:

459

Proc. 8th Int. Conf. on Applications of Statistics and Probability in Civil Engineering, Sydney; 1999. p. 415–22.

460

(21)

[6] Lloret-Cabot M, Fenton GA, Hicks MA. On the estimation of scale of fluctuation in geostatistics. Georisk:

461

Assessment and Management of Risk for Engineered Systems and Geohazards 2014; 8(2): 129–40.

462

[7] Lundberg AB, Li Y. Probabilistic characterization of a soft Scandinavian clay supporting a light quay

463

structure. In: Proc. 5th International Symposium on Geotechnical Safety and Risk, Rotterdam; 2015. p.170–75.

464

[8] Phoon KK, Kulhawy FH. Characterization of geotechnical variability. Canadian Geotechnical Journal 1999;

465

36 (4): 612–24.

466

[9] Uzielli M, Vannucchi G, Phoon KK. Random field characterisation of stress-normalised cone penetration

467

testing parameters. Géotechnique 2005; 55(1): 3–20.

468

[10] Zhao HF, Zhang LM, Xu Y, Chang D.S. Variability of geotechnical properties of a fresh landslide soil

469

deposit. Engineering Geology 2013; 166: 1–10.

470

[11] Fenton GA, Griffiths DV. Three-dimensional probabilistic foundation settlement. ASCE J. Geotech.

471

Geoenv. Eng. 2005; 131(2): 232–39.

472

[12] Jaksa MB, Goldsworthy JS, Fenton GA, Kaggwa WS, Griffiths DV, Kuo YL, Poulos HG. Towards reliable

473

and effective site investigations. Géotechnique 2005; 55(2): 109–21.

474

[13] Kuo YL, Jaksa MB, Kaggwa WS, Fenton GA, Griffiths DV, Goldsworthy JS. Probabilistic analysis of

475

multi-layered soil effects on shallow foundation settlement. In: Proc. 9th Australia New Zealand Conference on

476

Geomechanics, Aukland, New Zealand; 2004. p. 541–47.

477

[14] Griffiths DV, Fenton GA. Observations on two- and three-dimensional seepage through a spatially random

478

soil. In: Proc. 7th Int. Conf. on Applications of Statistics and Probability in Civil Engineering, Paris, France;

479

1995. p. 65–70.

480

[15] Griffiths DV, Fenton GA. Three-dimensional seepage through spatially random soil. ASCE J. Geotech.

481

Geoenv. Eng.1997; 123(2): 153–60.

482

[16] Griffiths DV, Fenton GA. Probabilistic analysis of exit gradients due to steady seepage. ASCE J. Geotech.

483

Geoenv. Eng. 1998; 124(9): 789–97.

484

[17] Popescu R, Prevost JH, Deodatis G. 3D effects in seismic liquefaction of stochastically variable soil

485

deposits. Géotechnique 2005; 55(1): 21–31.

486

[18] Griffiths DV, Huang J, Fenton GA. On the reliability of earth slopes in three dimensions. Proc. R. Soc. A

487

2009; 465(2110): 3145–64.

488

[19] Hicks MA, Spencer WA. Influence of heterogeneity on the reliability and failure of a long 3D slope.

489

Comput Geotech 2010; 37: 948–55.

490

(22)

[20] Hicks MA, Chen J, Spencer WA. Influence of spatial variability on 3D slope failures. In: Proc. 6th Int. Conf.

491

Computer Simulation Risk Analysis and Hazard Mitigation, Kefalonia; 2008. p. 335–42.

492

[21] Hicks MA, Nuttall JD, Chen J. Influence of heterogeneity on 3D slope reliability and failure consequence.

493

Computers and Geotechnics 2014; 61: 198–208.

494

[22] Li Y, Hicks MA. Comparative study of embankment reliability in three dimensions, In: Proc. 8th European

495

Conference on Numerical Methods in Geotechnical Engineering (NUMGE), Delft; 2014. p. 467–72.

496

[23] Li Y, Hicks MA, Nuttall JD. Probabilistic analysis of a benchmark problem for slope stability in 3D. In:

497

Proc. 3rd Int. Symp. Computational Geomech, Krakow, Poland; 2013. p. 641–8.

498

[24] Li YJ, Hicks MA, Nuttall, JD. Comparative analyses of slope reliability in 3D.Engineering Geology

499

2015;196: 12–23.

500

[25] Spencer WA. Parallel stochastic and finite element modeling of clay slope stability in 3D. PhD thesis,

501

University of Manchester, UK; 2007.

502

[26] Spencer WA, Hicks MA. 3D stochastic modelling of long soil slopes, In: Proc. 14th Conf. of Assoc. for

503

Computational Mechanics in Engineering, Belfast; 2006. p. 119–22.

504

[27] Spencer WA, Hicks MA. A 3D finite element study of slope reliability. In: Proc. 10th Int. Symp. Num

505

Models in Geomech, Rhodes; 2007. p. 539–43.

506

[28] Fenton GA, Griffiths DV. Risk assessment in geotechnical engineering. New York: John Wiley & Sons;

507

2008.

508

[29] Chiles JP, Delfiner P. Geostatistics: modeling spatial uncertainty. John Wiley & Sons; 2009.

509

[30] Lloret-Cabot M, Hicks MA, van den Eijnden AP. Investigation of the reduction in uncertainty due to soil

510

variability when conditioning a random field using Kriging. Géotechnique letters 2012; 2: 123–7.

511

[31] van den Eijnden AP, Hicks MA. Conditional simulation for characterizing the spatial variability of sand

512

state. In: Proc. 2nd International Symposium on Computational Geomechanics, Dubrovnik, Rhodes, Greece;

513

2011. p. 288–96.

514

[32] Vanmarcke EH, Fenton GA. Conditioned simulation of local fields of earthquake ground motion. Structural

515

Safety 1991; 10(1): 247–64.

516

[33] Journel, AG. Geostatistics for conditional simulation of ore bodies. Economic Geology 1974;69(5): 673–87.

517

[34] Olea, RA. Systematic sampling of spatial functions. Series of Spatial Analysis No. 7, Kansas Geological

518

Survey, Lawrence, Kans; 1984.

519

(23)

[35] Frimpong, S, Achireko PK. Conditional LAS stochastic simulation of regionalized variables in random

520

fields. Computational Geosciences 1998; 2: 37–45.

521

[36] Journel AG, Huijbregts CJ. Mining Geostatistics. New York: Academic Press; 1978.

522

[37] Fenton GA. Error evaluation of three random field generators. ASCE J. Eng. Mech. 1994; 120 (12): 2478–

523

97.

524

[38] Ji J, Liao HJ, Low BK. Modeling 2D spatial variation in slope reliability analysis using interpolated

525

autocorrelations. Computers and Geotechnics 2012; 40: 135–46.

526

[39] Phoon KK, Huang SP, Quek ST. Simulation of second-order processes using Karhunen–Loeve expansion.

527

Computers & Structures 2002; 80(12): 1049–60.

528

[40] Fenton GA, Vanmarcke EH. Simulation of random fields via local average subdivision. ASCE Journal of

529

Engineering Mechanics 1990; 116(8): 1733–49.

530

[41] Krige DG. A statistical approach to some basic mine valuation problems on the Witwatersrand. Journal of

531

the Chemical, Metallurgical and Mining Society of South Africa 1951; 52 (6): 119–39.

532

[42] Cressie N. The origins of Kriging. Mathematical Geology 1990; 22 (3): 239–52.

533

[43] Wackernagel H. Multivariate geostatistics: An introduction with applications. Germany: Springer; 2003.

534

[44] Fenton GA. Simulation and analysis of random fields. PhD Thesis, Princeton University; 1990.

535

[45] Vanmarcke EH. Random fields: analysis and synthesis. Cambridge, Massachusetts: The MIT Press; 1983.

536

[46] Fenton GA. Data analysis/geostatistics. In: Probabilistic methods in geotechnical engineering. Griffiths DV,

537

Fenton GA, editors. New York: Springer; 2007. p. 51–73.

538

[47] Smith IM, Griffiths DV, Margetts L. Programming the finite element method, 5th ed. New York: John

539

Wiley & Sons; 2013.

540

[48] Hicks MA, Samy K. Influence of heterogeneity on undrained clay slope stability. Quarterly Journal of

541

Engineering Geology and Hydrogeology 2002; 35: 41–9.

542

[49] Phoon KK, Kulhawy FH. Evaluation of geotechnical property variability. Canadian Geotechnical Journal

543

1999; 36(4): 625–39.

544

[50] European Committee for Standardisation (CEN). Eurocode 7: geotechnical design. Part 1: general rules. EN

545

1997-1, CEN; 2004.

546

[51] Hicks MA, Nuttall JD. Influence of soil heterogeneity on geotechnical performance and uncertainty: a

547

stochastic view on EC7. In: Proc. 10th International Probabilistic Workshop, Stuttgart; 2012. p. 215–27.

548 549

(24)

550 551 552 553 554

Appendix

555

A.1 Forming the left-hand-side matrix for Kriging

556

Suppose there are k×m CPT locations that follow a rectangular grid at the ground surface. That is, 557

there are k rows in the x direction and, within each row, m CPT profiles in the y direction (Figure 1). 558

Assuming that there are n data points for each CPT profile, the global numbering scheme for all the 559

CPT data points is shown in Figure A.1 for the case of k = 2. 560

Following the basic equation (6), of size N + 1 = k×m×n + 1, the left-hand-side matrix is formulated 561 as 562 1,1 1,2 1,3 1, 1, 1 1, 2 1, 3 1,2 2,1 2,2 2,3 2, 2, 1 2, 2 2, 3 2,2 3,1 3,2 3,3 3, 3, 1 3, 2 3, 3 3,2 ,1 ,2 ,3 , , 1 , 2 , 3 ,2

1

1

1

1

1

1

1

m m m m m m m m m m m m m m m lhs km km km km m km m km m km m km m + + + + + + + + + + + +

=

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

γ

v

v

v

v

v

v

v

v

2

2

2

2

2

2

2

2

2

2

2

2

2

1,( 1) 1 1,( 1) 2 1,( 1) 3 1, 2,( 1) 1 2,( 1) 2 2,( 1) 3 2, 3,( 1) 1 3,( 1) 2 3,( 1) 3 3, ,( 1) 1 ,( 1) 2 ,( 1) 3 ,

1

1

1

1

1

1

1

1

1

0

k m k m k m km k m k m k m km k m k m k m km km k m km k m km k m km km − + − + − + − + − + − + − + − + − + − + − + − +





v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

v

2

2

2

2

2

2

2

(A1) 563

in which

v

i j, is a matrix representing the correlation structure between CPTi and CPTj (where each 564

CPT has n data points), 565

(25)

( 1) 1,( 1) 1 ( 1) 1,( 1) 2 ( 1) 1,( 1) 3 ( 1) 1,( 1) ( 1) 2,( 1) 1 ( 1) 2,( 1) 2 ( 1) 2,( 1) 3 ( 1) 2,( 1) ( 1) 3,( 1) 1 ( 1) 3,( 1) 2 ( 1) 3,( , i n j n i n j n i n j n i n j n n i n j n i n j n i n j n i n j n n i n j n i n j n i n j i j

d

d

d

d

d

d

d

d

d

d

d

− + − + − + − + − + − + − + − + − + − + − + − + − + − + − + − + − + − + − + − + − + −

=

v

2

2

1) 3 ( 1) 3,( 1) ( 1) ,( 1) 1 ( 1) ,( 1) 2 ( 1) ,( 1) 3 ( 1) ,( 1) n i n j n n i n n j n i n n j n i n n j n i n n j n n

d

d

d

d

d

+ − + − + − + − + − + − + − + − + − + − +

2

2

(A2) 566

where (i, j) = 1, 2, 3, …, m, m+1, m+2, m+3, …, 2m, ……, (k-1)m+1, (k-1)m+2, (k-1)m+3, …, km and 567

dr,s (r = (i-1)n+1, …, (i-1)n+n) (s = (j-1)n+1, …, (j-1)n+n) are the components of the submatrix

v

i j, , 568

which can be expressed in the form of a covariance function between data points r and s (equation (2)). 569

A.2 Forming the right-hand-side vector for Kriging

570

The right-hand-side vector is formulated as 571 1 2 3

1

rhs km

= 

v

v

v

γ

v

(A3) 572

in which

v

pis a vector representing the correlation structure between the estimation point and CPTp, 573 ( 1) 1 ( 1) 2 ( 1) 3 ( 1) p n p n p n p p n n

d

d

d

d

− + − + − + − +

=

v

(A4) 574 where p = 1, 2, 3, …, m, m+1, m+2, m+3, …, 2m, ……, (k-1)m+1, (k-1)m+2, (k-1)m+3, …, km and dt 575

(t = (p-1)n+1, …, (p-1)n+n) are the components of the subvector

v

p, which can be expressed in the 576

form of a covariance function (equation (2)) between data points t and the point at which the value is 577

to be estimated (Figure A.1). 578

The unknown weight vector is 579

(26)

1 2 3 x km

µ

= 

λ

λ

λ

λ

λ

(A5) 580

in which

λ

qis the weight subvector for CPTq, 581 ( 1) 1 ( 1) 2 ( 1) 3 ( 1) q n q n q n q q n n

λ

λ

λ

λ

− + − + − + − +

=

λ

(A6) 582 where q = 1, 2, 3, …, m, m+1, m+2, m+3, …, 2m, ……, (k-1)m+1, (k-1)m+2, (k-1)m+3, …, km. 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 25

(27)

List of Figures

599

600

Figure 1. Example CPT sampling strategy (k = 2, m = 5) 601

Figure 2. Example illustrations of the unconditional random field (a), the Kriged field based on the 602

randomly simulated data (b), the Kriged field based on the CPT data (c), the conditional random field 603

(d), cross-sections (e and f) in the longitudinal direction taken from the Kriged field (c) and from the 604

conditional random field (d), respectively. Dashed circle indicates the position of the first CPT in 605

subfigures (a) and (c-d) 606

Figure 3. Vertical and horizontal covariance functions averaged over 200 realisations (θv = 1.0 m, θh = 607

3.0 m) 608

Figure 4. Flowchart for conditional RFEM simulation 609

Figure 5. Finite element mesh and possible numbered CPT locations at a cross-section through the 610

proposed 50 m long slope (dashed lines indicate the excavated soil mass and numbers correspond to 611

Gauss point locations within the finite elements) 612

Figure 6. Simulation results for Example 1 (based on θv = 1.0 m and 500 realisations per simulation) 613

Figure 7. Sampling efficiency indices for various values of

ξ

614

Figure 8. Kriging variance for various values of

ξ

(y–z slice at i = 5) 615

Figure 9. CPT layout illustration (plan view) for a single row (a) and two rows (b) 616

Figure 10. Influence of CPT location j during the second phase of site investigation (based on θv = 1.0 617

m and 500 realisations per simulation) 618

Figure 11. CPT layouts (plan views) for various numbers of boreholes (ncpt = 3, 5, 9, 17, 25 and Δ 619

denotes the distance between CPTs) 620

Cytaty

Powiązane dokumenty

a) Czy zaprezentowane w trakcie laboratorium metody faktycznie mogą stanowić samodzielne sposoby kompozycji, czy służyć za źródło inspiracji dla kompozytora? Odpowiedź uzasadnij.

Uzyskane wyniki przetwarzania pionowego profilowania sejsmicznego PPS 3C dla otworu W-1 w postaci transforma- cji PPS-WPG dla fal podłużnych i poprzecznych umożliwia- ją

A qualified electronic time stamp has a presumption of accuracy of the date and time it refers to, and integrity of the electronic data this date and time are associated with

10 цього закону містить перелік повноважень Національної поліції України у сфері запо- бігання та протидії домашньому насильству, а

Для проведення судової психолого-психіа- тричної експертизи (встановлення того, чи є підозрюваний симулянтом) необхідно зібрати список докумен- тів

Zespół pracowników naukowych Instytutu Nauk Biblijnych (stan osobowy na dzien´ 15.06.2003 r.) tworzyli: prof.. Ryszard Rubinkiewicz SDB – dyrektor

Komentarz Mounce’a – wbrew wspomnianemu powszechnemu trendowi – stanowi przyczynek staraj ˛ acy sie˛ wykazac´, z˙e wszystkie trzy Listy (1-2 Tm i Tt) zostały napisane

Prowadzenie dziennika jest ciągłym powracaniem do źródeł tworzenia. W przypadku Nałkowskiej wyraża nieumiejętność bezpośredniego przeżywania, by­ cia